1 Introduction

The purpose of this short paper is to refine a classical theorem of Bebutov [2] and Kakutani [4] on dynamical systems. We call (XT) a flow if X is a compact metric space and

$$\begin{aligned} T:\mathbb {R}\times X\rightarrow X, \quad (t, x) \mapsto T_t x \end{aligned}$$

is a continuous action of \(\mathbb {R}\). We define \(\mathrm {Fix}(X,T)\) (sometimes abbreviated to \(\mathrm {Fix}(X)\)) as the set of \(x\in X\) satisfying \(T_t x=x\) for all \(t\in \mathbb {R}\). We define \(C(\mathbb {R})\) as the space of continuous maps \(\varphi : \mathbb {R}\rightarrow [0,1]\). It is endowed with the topology of uniform convergence over compact subsets of \(\mathbb {R}\), namely the topology given by the distance

$$\begin{aligned} \sum _{n=1}^\infty 2^{-n} \max _{|t|\le n} |\varphi (t)-\psi (t)|, \quad (\varphi , \psi \in C(\mathbb {R})). \end{aligned}$$
(1.1)

The group \(\mathbb {R}\) continuously acts on it by the translation:

$$\begin{aligned} \mathbb {R}\times C(\mathbb {R}) \rightarrow C(\mathbb {R}), \quad (s, \varphi (t))\mapsto \varphi (t+s). \end{aligned}$$
(1.2)

A continuous map \(f:X\rightarrow C(\mathbb {R})\) is called an embedding of a flow (XT) if f is an \(\mathbb {R}\)-equivariant topological embedding. Bebutov [2] and Kakutani [4] found that the \(\mathbb {R}\)-action on \(C(\mathbb {R})\) has the following remarkable “universality”:

Theorem 1.1

(Bebutov–Kakutani) A flow (XT) can be equivariantly embedded in \(C(\mathbb {R})\) if and only if \(\mathrm {Fix}(X,T)\) can be topologically embedded in the unit interval [0, 1].

The “only if” part is trivial because the set of fixed points of \(C(\mathbb {R})\) is homeomorphic to [0, 1]. So the main statement is the “if” part.

Although the Bebutov–Kakutani theorem is clearly a nice theorem, it has one drawback: The space \(C(\mathbb {R})\) is not compact (nor locally compact). So it is not a “flow” in the above definition. This poses the following problem:

Problem 1.2

Is there a compact invariant subset of \(C(\mathbb {R})\) satisfying the same universality?

The purpose of this paper is to solve this problem affirmatively. Let \(L(\mathbb {R})\) be the set of maps \(\varphi :\mathbb {R}\rightarrow [0,1]\) satisfying the one-Lipschitz condition:

$$\begin{aligned} \forall s, t\in \mathbb {R}: \quad |\varphi (s)-\varphi (t)| \le |s-t|. \end{aligned}$$
(1.3)

\(L(\mathbb {R})\) is a subset of \(C(\mathbb {R})\). It is compact with respect to the distance (1.1) by Ascoli–Arzela’s theorem. The \(\mathbb {R}\)-action (1.2) preserves \(L(\mathbb {R})\). So it becomes a flow. Our main result is the following. This solves [3, Question 4.1].

Theorem 1.3

A flow (XT) can be equivariantly embedded in \(L(\mathbb {R})\) if and only if \(\mathrm {Fix}(X,T)\) can be topologically embedded in the unit interval [0, 1].

As in the case of the Bebutov–Kakutani theorem, the “only if” part is trivial because the fixed point set \(\mathrm {Fix}(L(\mathbb {R}))\) is homeomorphic to [0, 1]. Since \(L(\mathbb {R})\) is compact, it is a more reasonable choice of such a “universal flow”.

The proof of Theorem 1.3 is based on the techniques originally used in the proof of the Bebutov–Kakutani theorem (in particular, the idea of local section). A main new ingredient is the topological argument given in Sect. 2, which has some combinatorial flavor.

Remark 1.4

Problem 1.2 asks us to find a universal flow smaller than\(C(\mathbb {R})\). If we look for a universal flow larger than\(C(\mathbb {R})\), then it is much easier to find an example. Let \(L^\infty (\mathbb {R})\) be the set of \(L^\infty \)-functions \(\varphi : \mathbb {R}\rightarrow [0,1]\). (We identify two functions which are equal to each other almost everywhere.) We consider the weak\(^*\) topology on it. Namely a sequence \(\{\varphi _n\}\) in \(L^\infty (\mathbb {R})\) converges to \(\varphi \in L^\infty (\mathbb {R})\) if for every \(L^1\)-function \(\psi :\mathbb {R}\rightarrow \mathbb {R}\)

$$\begin{aligned} \lim _{n\rightarrow \infty } \int _{\mathbb {R}} \varphi _n(t) \psi (t) \, dt = \int _{\mathbb {R}} \varphi (t) \psi (t) \, dt. \end{aligned}$$

Then \(L^\infty (\mathbb {R})\) is compact and metrizable by Banach–Alaoglu’s theorem and the separability of the space of \(L^1\)-functions, respectively. The group \(\mathbb {R}\) acts continuously on it by translation. So it becomes a flow. Note that \(\mathrm {Fix}\left( L^\infty (\mathbb {R})\right) \) is homeomorphic to [0, 1] and that the natural inclusion map \(C(\mathbb {R}) \subset L^\infty (\mathbb {R})\) is an equivariant continuous injection. Then the Bebutov–Kakutani theorem implies the universality of \(L^\infty (\mathbb {R})\): A flow (XT) can be equivariantly embedded in \(L^\infty (\mathbb {R})\) if and only if \(\mathrm {Fix}(X,T)\) can be topologically embedded in [0, 1].

2 Topological Preparations

Let a be a positive number. We define L[0, a] as the space of maps \(\varphi :[0,a]\rightarrow [0,1]\) satisfying

$$\begin{aligned} \forall s, t\in [0,a]: \quad |\varphi (s)-\varphi (t)| \le |s-t|. \end{aligned}$$

L[0, a] is endowed with the distance \(\left| \left| \varphi -\psi \right| \right| _\infty = \max _{0 \le t\le a} |\varphi (t)-\psi (t)|\). We define \(F_L[0, a] \subset L[0,a]\) as the space of constant functions \(\varphi :[0,a]\rightarrow [0,1]\), which is homeomorphic to [0, 1].

Let (Xd) be a compact metric space. We define \(C\left( X,L[0,a]\right) \) as the space of continuous maps \(f:X\rightarrow L[0,a]\), which is endowed with the distance

$$\begin{aligned} \max _{x\in X} \left| \left| f(x)-g(x)\right| \right| _\infty . \end{aligned}$$

Lemma 2.1

Let \(f\in C\left( X, L[0,a]\right) \) and suppose there exists \(0<\tau <1\) satisfying

$$\begin{aligned} \forall x\in X, \forall s, t\in [0,a]: \quad |f(x)(s)-f(x)(t)| \le \tau |s-t|. \end{aligned}$$
(2.1)

Then for any \(\delta >0\) there exists \(g\in C\left( X, L[0,a]\right) \) satisfying

  1. (1)

    \(\max _{x\in X} \left| \left| f(x)-g(x)\right| \right| _\infty <\delta \).

  2. (2)

    \(g(x)(0)=f(x)(0)\) and \(g(x)(a)=f(x)(a)\) for all \(x\in X\).

  3. (3)

    \(g(X)\cap F_L[0,a] = \emptyset \).

Proof

We take \(0<b<c<a\) satisfying \(b=a-c < \delta /4\). We take an open covering \(\{U_1, \ldots ,U_M\}\) of X satisfying

$$\begin{aligned} \forall 1\le m\le M: \quad \mathrm {diam}f(U_m) < \min \left( \frac{\delta }{4}, \frac{(1-\tau )b}{2}\right) . \end{aligned}$$
(2.2)

We take a point \(p_m\in U_m\) for each m. We choose a natural number N satisfying

$$\begin{aligned} N > M, \quad \Delta \overset{\mathrm {def}}{=} \frac{c-b}{N-1} < \frac{\delta }{4}. \end{aligned}$$

We divide the interval [bc] into \((N-1)\) intervals of length \(\Delta \):

$$\begin{aligned} b= a_1<a_2<\dots <a_{N}=c, \quad a_{n+1}-a_n = \Delta \quad (\forall 1\le n\le N-1). \end{aligned}$$

Set \(A= \{a_1, \ldots , a_N\}\) and define a vector \(e\in \mathbb {R}^A\) by \(e=(1,1, \ldots , 1)\). Notice that \(f(p_m)|_{A}\) is an element of \([0,1]^{A}\). Since \(N>M\) we can choose \(u_1, \ldots , u_M\in [0,1]^A\) satisfying

  1. (1)

    \(|f(p_m)(a_n) - u_m(a_n)| < \min \left( \delta /4, (1-\tau )b/2\right) \) for all \(1\le m\le M\) and \(1\le n\le N\).

  2. (2)

    \(|u_m(a_{n+1})-u_m(a_n)| < \Delta \) for all \(1\le m\le M\) and \(1\le n\le N-1\).

  3. (3)

    The \((M+1)\) vectors \(e, u_1, \ldots , u_M\) are linearly independent.

Let \(\{h_m\}_{m=1}^M\) be a partition of unity on X satisfying \(\mathrm {supp}\, h_m\subset U_m\) for all m. For \(x\in X\) we define a piecewise linear function \(g(x):[0,a]\rightarrow [0,1]\) as follows. (We set \(a_0=0\) and \(a_{N+1}=a\).)

  • \(g(x)(0) = f(x)(0)\) and \(g(x)(a)= f(x)(a)\).

  • \(g(x)(a_n) = \sum _{m=1}^M h_m(x) u_m(a_n)\) for \(1\le n\le N\).

  • We extend g(x) linearly. Namely, for \(t= (1-\lambda ) a_n + \lambda a_{n+1}\) with \(0\le \lambda \le 1\) and \(0\le n\le N\) we set \(g(x)(t) = (1-\lambda ) g(a_n) + \lambda g(a_{n+1})\).

\(\square \)

Claim 2.2

\(g(x)\in L[0,a]\) and \(\left| \left| g(x)-f(x)\right| \right| _\infty < \delta \).

Proof

For proving \(g(x)\in L[0,a]\) it is enough to show \(|g(x)(a_{n+1})-g(x)(a_n)| \le |a_{n+1}-a_n|\) for all \(0\le n\le N\). For \(1\le n\le N-1\), this is a direct consequence of the property (2) of \(u_m\). So we consider the case of \(n=0\). (The case of \(n=N\) is the same).

$$\begin{aligned} \begin{aligned} |g(x)(b) - f(x)(0)| \le&\sum _{m=1}^M h_m(x) |u_m(b)-f(p_m)(b)| + \sum _{m=1}^M h_m(x) |f(p_m)(b) - f(x)(b)| \\&+ |f(x)(b)-f(x)(0)|. \end{aligned} \end{aligned}$$

We apply to each term of the right-hand side the property (1) of \(u_m\), \(\mathrm {diam}f(U_m) < (1-\tau )b/2\) in (2.2) and \(|f(x)(b)-f(x)(0)| \le \tau b\) in (2.1) respectively. Then this is bounded by

$$\begin{aligned} \frac{(1-\tau )b}{2} + \frac{(1-\tau )b}{2} + \tau b = b. \end{aligned}$$

This proves \(g(x)\in L[0,a]\).

Next we show \(|g(x)(a_n) - f(x)(a_n)| < \delta /2\) for all \(0\le n\le N+1\). For \(n=0, N+1\), this is trivial. For \(1\le n\le N\), we can bound \(|g(x)(a_n)-f(x)(a_n)|\) from above by

$$\begin{aligned} \begin{aligned}&\sum _{m=1}^M h_m(x) |u_m(a_n)-f(p_m)(a_n)| + \sum _{m=1}^M h_m(x) |f(p_m)(a_n)-f(x)(a_n)| \\&\quad< \frac{\delta }{4}+ \frac{\delta }{4} = \frac{\delta }{2} \quad \left( \text {by the property (1) of}~u_m~\text {and}~\mathrm {diam}f(U_m) < \frac{\delta }{4}~\text {in}~(2.2)\right) . \end{aligned} \end{aligned}$$

Finally, let \(a_n<t<a_{n+1}\). We can bound \(|g(x)(t)-f(x)(t)|\) by

$$\begin{aligned} \begin{aligned}&|g(x)(t)-g(x)(a_n)| + |g(x)(a_n)-f(x)(a_n)| + |f(x)(a_n)-f(x)(t)| \\&\quad< 2(a_{n+1}-a_n) + \frac{\delta }{2} \quad (\text {by}~f(x), g(x)\in L[0,a]) \\&\quad< \delta \quad \left( \text {by}~a_{n+1}-a_n \le \max (b, \Delta ) < \frac{\delta }{4}\right) . \end{aligned} \end{aligned}$$

\(\square \)

For every \(x\in X\), the function \(g(x):[0,a]\rightarrow [0,1]\) is a non-constant function because

$$\begin{aligned} g(x)|_{\Lambda } = \sum _{m=1}^M h_m(x) u_m \not \in \mathbb {R} e \quad (\text {by the property (3) of}~u_m). \end{aligned}$$

This proves the statement. \(\square \)

We need two lemmas on linear algebra. For \(u=(x_1, \ldots , x_{n+1})\in \mathbb {R}^{n+1}\) we set

$$\begin{aligned} Du = (x_2-x_1,x_3-x_2, \ldots , x_{n+1}-x_{n}) \in \mathbb {R}^{n}. \end{aligned}$$

Lemma 2.3

Let \(l\ge m+1\) and set \(e=(\underbrace{1,1, \ldots , 1}_l) \in \mathbb {R}^{l}\). The set of \((u_1, \ldots , u_m) \in \mathbb {R}^{l+1}\times \dots \times \mathbb {R}^{l+1} = \left( \mathbb {R}^{l+1}\right) ^m\) such that

$$\begin{aligned} \text {the vectors}~e, Du_1, Du_2, \ldots , Du_{m}~\text { are linearly independent} \end{aligned}$$
(2.3)

is open and dense in \(\left( \mathbb {R}^{l+1}\right) ^{m}\).

Proof

The condition (2.3) defines a Zariski open set in \(\left( \mathbb {R}^{l+1}\right) ^{m}\). So it is enough to show that the set is non-empty because a non-empty Zariski open set is always dense in the Euclidean topology. We set

$$\begin{aligned} u_i = (\underbrace{-1, \ldots , -1}_{i}, \, \underbrace{0, \ldots , 0}_{l+1-i}) , \quad (1\le i\le m). \end{aligned}$$

Then

$$\begin{aligned} Du_i = (\underbrace{0, \ldots , 0}_{i-1},\, 1, \, \underbrace{0, \ldots , 0}_{l-i}). \end{aligned}$$

The vectors \(e, Du_1, \ldots Du_{m}\) are linearly independent. \(\square \)

Lemma 2.4

Let \(n > l \ge 2m\). The set of \((u_1, \ldots , u_m) \in \mathbb {R}^n\times \dots \times \mathbb {R}^n = \left( \mathbb {R}^n\right) ^m\) such that, for any integer \(\alpha \) with \(2\le \alpha \le n-l+1\),

$$\begin{aligned} u_1|_1^l ,\> u_1|_{\alpha }^{\alpha +l-1} ,\> u_2|_1^l, \> u_2|_{\alpha }^{\alpha +l-1}, \ldots , u_m|_{1}^l, \> u_m|_{\alpha }^{\alpha +l-1}~ \text {are linearly independent in}~\mathbb {R}^l \end{aligned}$$
(2.4)

is open and dense in \(\left( \mathbb {R}^n\right) ^m\). Here for \(u_i = (x_{i1}, \ldots , x_{i n})\)

$$\begin{aligned} u_i|_{1}^l = (x_{i 1}, \ldots , x_{i l}), \quad u_i|_\alpha ^{\alpha +l-1} = (x_{i, \alpha }, \ldots , x_{i, \alpha + l-1}). \end{aligned}$$

Proof

The defined set \(A=\cap _{\alpha =2}^{n-l+1}A_\alpha \), where

$$\begin{aligned} A_\alpha =\{(u_1, \ldots ,u_m)\in \left( \mathbb {R}^n\right) ^m|\, (2.4)~\text {is satisfied}\}, \end{aligned}$$

is a Zariski open set in \(\left( \mathbb {R}^n\right) ^m\). Hence it is enough to show that every \(A_\alpha \) is nonempty. For each fixed \(2 \le \alpha \le n-l+1\), we define \(u_i = (x_{i1}, \ldots , x_{i n})\)\((1 \le i\le m)\) by

$$\begin{aligned} x_{ij} = 1 \quad \left( j= i, \alpha + l-i \right) , \quad x_{ij} = 0 \quad (\text {otherwise}). \end{aligned}$$

Then it is direct to check that \((u_1, \ldots ,u_m)\in A_\alpha \). One can also use a proof from [5, Lemma 5.5]. \(\square \)

Lemma 2.5

Let \(f\in C\left( X, L[0,a]\right) \) and suppose there exists \(0<\tau <1\) satisfying (2.1). Then for any \(\delta >0\) there exists \(g\in C\left( X, L[0,a]\right) \) satisfying

  1. (1)

    \(\max _{x\in X} \left| \left| f(x)-g(x)\right| \right| _\infty < \delta \).

  2. (2)

    \(g(x)(0)=f(x)(0)\) and \(g(x)(a)=f(x)(a)\) for all \(x\in X\).

  3. (3)

    If \(x,y\in X\) and \(0\le \varepsilon \le a/2\) satisfy

    $$\begin{aligned} \forall t\in [0,a-\varepsilon ]: g(x)(t+\varepsilon ) = g(y)(t) \end{aligned}$$

    then \(\varepsilon =0\) and \(d(x,y) < \delta \).

Proof

Except for the use of the above two lemmas on linear algebra, the proof is close to Lemma 2.1. We take \(0<b<c<a\) with \(b=a-c < \min \left( \delta /4, a/4\right) \). We take an open covering \(\{U_1, \ldots ,U_M\}\) satisfying \(\mathrm {diam}\, U_m < \delta \) and \(\mathrm {diam}f(U_m) < \min \left( \delta /4, (1-\tau )b/2\right) \) for all \(1\le m\le M\). Take \(p_m\in U_m\) for each m. Let \(N\ge 2\) be a natural number and set \(\Delta = (c-b)/(N-1)\). We introduce a partition \(b=a_1<a_2<\dots <a_N=c\) by \(a_n = b + (n-1)\Delta \). We set \(A= \{a_1, \ldots , a_N\}\) and \(\Lambda = A\cap [b,a/4] = \{a_1, \ldots , a_L\}\). We also set \(e=(\underbrace{1,1, \ldots , 1}_L)\in \mathbb {R}^{L}\). We choose N sufficiently large so that

$$\begin{aligned} \Delta < \frac{\delta }{4}, \quad N> L \ge 2M. \end{aligned}$$

Since \(L\ge 2M\ge M+1\), by using Lemmas 2.3 and 2.4, we can choose \(u_1, \ldots , u_M\in [0,1]^A\) satisfying

  1. (1)

    \(|f(p_m)(a_n) - u_m(a_n)| < \min \left( \delta /4, (1-\tau )b/2\right) \) for all \(1\le m\le M\) and \(1\le n\le N\).

  2. (2)

    \(|u_m(a_{n+1})-u_m(a_n)| < \Delta \) for all \(1\le m\le M\) and \(1\le n\le N-1\).

  3. (3)

    Define \(D_L u_m = (u_m(a_2)-u_m(a_1), \ldots , u_m(a_{L+1})-u_m(a_L))\in \mathbb {R}^L\). Then the \((M+1)\) vectors \(e, D_L u_1, \ldots , D_L u_M\) in \(\mathbb {R}^L\) are linearly independent.

  4. (4)

    For any \(\varepsilon > 0\) with \(\varepsilon + \Lambda \subset A\),

    $$\begin{aligned} u_1|_{\Lambda },\> u_1|_{\varepsilon +\Lambda },\> u_2|_{\Lambda },\> u_2|_{\varepsilon +\Lambda }, \ldots , u_m|_{\Lambda }, \> u_m|_{\varepsilon +\Lambda }~\text {are linearly independent in}~ \mathbb {R}^\Lambda \end{aligned}$$

For \(x\in X\) we define \(g(x):[0,a]\rightarrow [0,1]\) in the same way as in the proof of Lemma 2.1. Namely, we set \(g(x)(0)= f(x)(0)\), \(g(x)(a)= f(x)(a)\) and \(g(x)(a_n) = \sum _{m=1}^M h_m(x) u_m(a_n)\) for \(1\le n\le N\), where \(\{h_m\}\) is a partition of unity satisfying \(\mathrm {supp}\, h_m\subset U_m\). We extend g(x) to [0, a] by linearity. It follows that \(g(x)\in L[0,a]\) and \(\left| \left| g(x)-f(x)\right| \right| _\infty < \delta \) as before. We need to check the property (3) of the statement. Suppose there exist \(x,y\in X\) and \(0\le \varepsilon \le a/2\) satisfying \(g(x)(t+\varepsilon ) = g(y)(t)\) for all \(0\le t\le a-\varepsilon \).

First we show \(\varepsilon + \Lambda \subset A\). Otherwise, \((\varepsilon +\Lambda ) \cap A = \emptyset \). Then it follows from the piecewise linearity that the function g(y)(t) becomes differentiable at every \(t\in \Lambda \), which implies

$$\begin{aligned} g(y)(a_{n+1}) - g(y)(a_n) = g(y)(a_{n+2})-g(y)(a_{n+1}) \quad (1\le n\le L-1), \end{aligned}$$

and hence

$$\begin{aligned} \sum _{m=1}^M h_m(y) \left( u_m(a_{n+1}) - u_m(a_n)\right) = \sum _{m=1}^M h_m(y)\left( u_m(a_{n+2})-u_m(a_{n+1})\right) \quad (1\le n\le L-1). \end{aligned}$$

This means that \(\sum _{m=1}^M h_m(y) D_L u_m \in \mathbb {R} e\), which contradicts the property (3) of \(u_m\). So we must have \(\varepsilon +\Lambda \subset A\).

The equation \(g(x)(t+\varepsilon ) = g(y)(t)\)\((0\le t\le a-\varepsilon )\) implies

$$\begin{aligned} \sum _{m=1}^M h_m(x) u_m|_{\varepsilon +\Lambda } = \sum _{m=1}^M h_m(y) u_m|_{\Lambda }. \end{aligned}$$

It follows from the property (4) of \(u_m\) that \(\varepsilon =0\) and \(h_m(x)=h_m(y)\) for all \(1\le m\le M\). Then \(x,y\in U_m\) for some m and hence \(d(x,y) \le \mathrm {diam}\, U_m < \delta \). \(\square \)

3 Proof of Theorem 1.3

Let (XT) be a flow. Set \(F=\mathrm {Fix}(X,T)\). We define \(F_L = \mathrm {Fix}\left( L(\mathbb {R})\right) \). Namely \(F_L\) is the space of constant maps \(\varphi :\mathbb {R}\rightarrow [0,1]\), which is homeomorphic to [0, 1]. Suppose there exists a topological embedding \(h:F\rightarrow F_L\). We would like to show that there exists an equivariant embedding \(f:X\rightarrow L(\mathbb {R})\) with \(f|_F=h\). We define \(C_{T,h}\left( X,L(\mathbb {R})\right) \) as the space of equivariant continuous maps \(f:X\rightarrow L(\mathbb {R})\) satisfying \(f|_{F} = h\), which is endowed with the compact-open topology. For \(f\in C_{T,h}\left( X, L(\mathbb {R})\right) \) we define \(\mathrm {Lip}(f)\) as the supremum of

$$\begin{aligned} \frac{|f(x)(t)-f(x)(s)|}{|s-t|} \end{aligned}$$

over all \(x\in X\) and \(s,t\in \mathbb {R}\) with \(s\ne t\).

Lemma 3.1

The space \(C_{T,h}\left( X, L(\mathbb {R})\right) \) is not empty. Moreover for any \(\delta >0\) there exists \(f\in C_{T,h}\left( X, L(\mathbb {R})\right) \) satisfying \(\mathrm {Lip}(f)\le \delta \).

Proof

Consider the map

$$\begin{aligned} F\ni x\rightarrow h(x)(0) \in [0,1]. \end{aligned}$$

By the Tietze extension theorem, we can extend this function to a continuous map \(h_0:X\rightarrow [0,1]\). Let \(\varphi : \mathbb {R}\rightarrow [0,1]\) be a smooth function satisfying

$$\begin{aligned} \int _{-\infty }^\infty \varphi (t)\, dt = 1, \quad \int _{-\infty }^\infty \varphi '(t) \, dt \le \min \left( 1, \delta \right) . \end{aligned}$$

For \(x\in X\) we define \(f(x): \mathbb {R} \rightarrow [0,1]\) by

$$\begin{aligned} f(x)(t) = \int _{-\infty }^\infty \varphi (t-s) h_0(T_s x) \, ds. \end{aligned}$$

Then \(|f(x)'(t)|\le \min (1,\delta )\) and \(f=h\) on F. Hence \(f\in C_{T,h}\left( X, L(\mathbb {R})\right) \) and \(\mathrm {Lip}(f)\le \delta \). \(\square \)

We borrow the next lemma from Auslander [1, p. 186, Corollary 6].

Lemma 3.2

Let \(p\in X{\setminus } F\). There exist \(a >0\) and a closed set \(S\subset X\) containing p such that the map

$$\begin{aligned} {[}-a, a] \times S\rightarrow X, \quad (t, x)\mapsto T_t x \end{aligned}$$
(3.1)

is a continuous injection whose image contains an open neighborhood of p in X. We call (aS) a local section around p and denote the image of (3.1) by \([-a,a] \cdot S\).

Proof

We explain the proof for the convenience of readers. We can find \(c<0\) and a continuous function \(h:X\rightarrow [0,1]\) satisfying \(T_c p \not \in \mathrm {supp}\, h\) and \(h=1\) on a neighborhood of p. We define \(f:X\rightarrow \mathbb {R}\) by

$$\begin{aligned} f(x) = \int _c^0 h(T_t x) dt. \end{aligned}$$

We choose \(0<a<|c|\) and a closed neighborhood A of p satisfying

$$\begin{aligned} \bigcup _{|t|\le a} T_t(A) \subset \{h=1\}, \quad \bigcup _{|t|\le a} T_{t+c}(A) \cap \mathrm {supp}\, h = \emptyset . \end{aligned}$$

It follows that \(f\left( T_t x\right) = f(x) + t\) for any \(x\in A\) and \(|t|\le a\). Set \(S= \{x\in A|\, f(x)=f(p)\}\). Then (aS) becomes a local section. Indeed if \(x,y\in S\) and \(s,t\in [-a,a]\) satisfy \(T_s x = T_t y\), then \(s+f(p) = f(T_s x)= f(T_t y) = t+f(p)\) and hence \(s=t\) and \(x=y\). Thus the map (3.1) is injective. We take \(0<b<a\) and an open neighborhood U of p satisfying \(\bigcup _{|t|<b} T_t(U) \subset A\). Then the set \([-a,a]\cdot S\) contains

$$\begin{aligned} \{x\in U|\, -b<f(x)-f(p)<b\} \end{aligned}$$
(3.2)

because if \(x\in U\) satisfies \(t \overset{\mathrm {def}}{=} f(x)-f(p) \in (-b,b)\) then \(f(T_{-t} x) = f(x)-t = f(p)\) (i.e. \(T_{-t}x\in S\)) and \(x = T_t(T_{-t} x) \in [-a,a]\cdot S\). The set (3.2) is an open neighborhood of p. \(\square \)

Lemma 3.3

For any point \(p\in X{\setminus } F\) there exists a closed neighborhood A of p in X such that the set

$$\begin{aligned} G(A) = \left\{ f\in C_{T,h}\left( X, L(\mathbb {R})\right) |\, f(A) \cap F_L = \emptyset \right\} \end{aligned}$$
(3.3)

is open and dense in the space \(C_{T,h}\left( X, L(\mathbb {R})\right) \).

Proof

Take a local section (aS) around p. For \(x\in X\) we define \(H(x)\subset \mathbb {R}\) (the set of hitting times) as the set of \(t\in \mathbb {R}\) satisfying \(T_t x\in S\). Any two distinct \(s,t\in H(x)\) satisfy \(|s-t|> a\). Notice that if \(x\in F\) then \(H(x)=\emptyset \). We denote by \(\mathrm {Int}\left( [-a,a]\cdot S\right) \) the interior of \([-a,a]\cdot S\). We choose a closed neighborhood \(A_0\) of p in S satisfying \(A_0\subset \mathrm {Int}\left( [-a,a]\cdot S\right) \). We define a closed neighborhood A of p in X by

$$\begin{aligned} A = \bigcup _{|t|\le a} T_t(A_0). \end{aligned}$$

We choose a continuous function \(q: S\rightarrow [0,1]\) satisfying \(q=1\) on \(A_0\) and \(\mathrm {supp}\, q \subset \mathrm {Int}\left( [-a,a]\cdot S\right) \).

The set G(A) defined in (3.3) is obviously open. So it is enough to prove that it is dense. Take \(f\in C_{T,h}\left( X, L(\mathbb {R})\right) \) and \(0<\delta <1\). By Lemma 3.1 we can find \(f_0\in C_{T,h}\left( X, L(\mathbb {R})\right) \) satisfying \(\mathrm {Lip}(f_0)\le 1/2\). We define \(f_1\in C_{T,h}\left( X, L(\mathbb {R})\right) \) by

$$\begin{aligned} f_1(x)(t) = (1-\delta ) f(x)(t) + \delta f_0(x)(t). \end{aligned}$$

It follows \(\mathrm {Lip}(f_1) \le 1-\delta /2 < 1\). We apply Lemma 2.1 to the map

$$\begin{aligned} X\ni x\mapsto f_1(x)|_{[0,a]} \in L[0,a]. \end{aligned}$$

Then we find \(g\in C\left( X, L[0,a]\right) \) satisfying

  1. (1)

    \(|g(x)(t)-f_1(x)(t)| < \delta \) for all \(x\in X\) and \(0\le t\le a\).

  2. (2)

    \(g(x)(0)= f_1(x)(0)\) and \(g(x)(a)= f_1(x)(a)\) for all \(x\in X\).

  3. (3)

    \(g(X) \cap F_L[0,a] = \emptyset \).

We set \(u(x)(t) = g(x)(t)-f_1(x)(t)\) for \(x\in X\) and \(0\le t\le a\). We define \(g_1\in C_{T,h}\left( X, L(\mathbb {R})\right) \) as follows: Let \(x\in X\).

  • For each \(s\in H(x)\), we set

    $$\begin{aligned} g_1(x)(t) = f_1(x)(t) + q(T_s x)\cdot u(T_s x)(t-s) \quad \text {for}~ t\in [s,s+a]. \end{aligned}$$
  • For \(t\in \mathbb {R}{\setminus } \bigcup _{s\in H(x)} [s,s+a]\), we set \(g_1(x)(t) = f_1(x)(t)\).

This satisfies

$$\begin{aligned} |g_1(x)(t)-f(x)(t)| \le |g_1(x)(t)-f_1(x)(t)| + |f_1(x)(t)-f(x)(t)| \le 3\delta \end{aligned}$$

for all \(x\in X\) and \(t\in \mathbb {R}\). If \(x\in A\) then there exists \(s\in [-a,a]\) with \(T_s x\in A_0\) and hence

$$\begin{aligned} g_1(x)(s+t) = g(T_s x)(t) \quad \text {for}~t\in [0,a]. \end{aligned}$$

It follows from the property (3) of g that the function \(g_1(x)\) is not constant. Thus \(g_1\in G(A)\). Since f and \(\delta \) are arbitrary, this proves that G(A) is dense in \(C_{T,h}\left( X, L(\mathbb {R})\right) \). \(\square \)

Lemma 3.4

For any two distinct points p and q in \(X{\setminus } F\) there exist closed neighborhoods B and C of p and q in X respectively such that the set

$$\begin{aligned} G(B,C) = \left\{ f\in C_{T,h}\left( X, L(\mathbb {R})\right) |\, f(B) \cap f(C) = \emptyset \right\} \end{aligned}$$
(3.4)

is open and dense in \(C_{T,h}\left( X, L(\mathbb {R})\right) \).

Proof

Take local sections \((a, S_1)\) and \((a, S_2)\) around p and q respectively. We can assume that \([-a,a]\cdot S_1\) and \([-a,a]\cdot S_2\) are disjoint with each other. For \(x\in X\) we define H(x) as the set of \(t\in \mathbb {R}\) satisfying \(T_t x\in S_1\cup S_2\). We choose closed neighborhoods \(B_0\) of p in \(S_1\) and \(C_0\) of q in \(S_2\) respectively satisfying \(B_0\subset \mathrm {Int}\left( [-a,a]\cdot S_1\right) \) and \(C_0\subset \mathrm {Int}\left( [-a,a]\cdot S_2\right) \). We take a continuous function \(\tilde{q}:X\rightarrow [0,1]\) satisfying \(\tilde{q}=1\) on \(B_0\cup C_0\) and \(\mathrm {supp}\, \tilde{q}\subset \mathrm {Int}\left( [-a,a]\cdot S_1\right) \cup \mathrm {Int}\left( [-a,a]\cdot S_2\right) \). We define closed neighborhoods B and C of p and q respectively by

$$\begin{aligned} B = \bigcup _{|t|\le a/4} T_t(B_0), \quad C = \bigcup _{|t|\le a/4} T_t(C_0). \end{aligned}$$

The set G(BC) defined in (3.4) is open. We show that it is dense. Take \(f\in C_{T,h}\left( X, L(\mathbb {R})\right) \) and \(0<\delta <1\). We can assume that

$$\begin{aligned} \delta < d(B_0,C_0) \overset{\mathrm {def}}{=} \min _{x\in B_0, y\in C_0} d(x,y). \end{aligned}$$
(3.5)

We define \(f_1\in C_{T,h}\left( X, L(\mathbb {R})\right) \) exactly in the same way as in the proof of Lemma 3.3. It satisfies \(\mathrm {Lip}(f_1) \le 1-\delta /2\) and \(|f(x)(t)-f_1(x)(t)| \le 2\delta \) for all \(x\in X\) and \(t\in \mathbb {R}\).

We apply Lemma 2.5 to the map

$$\begin{aligned} X\ni x\mapsto f_1(x)|_{[0,a]} \in L[0,a]. \end{aligned}$$

Then we find \(g\in C\left( X, L[0,a]\right) \) satisfying

  1. (1)

    \(|g(x)(t)- f_1(x)(t)| < \delta \) for all \(x\in X\) and \(0\le t \le a\).

  2. (2)

    \(g(x)(0)=f_1(x)(0)\) and \(g(x)(a)=f_1(x)(a)\) for all \(x\in X\).

  3. (3)

    If \(x,y\in X\) and \(0\le \varepsilon \le a/2\) satisfy

    $$\begin{aligned} \forall t\in [0,a-\varepsilon ]: g(x)(t+\varepsilon ) = g(y)(t) \end{aligned}$$

    then \(d(x,y) < \delta \).

We set \(u(x)(t) = g(x)(t)-f_1(x)(t)\) for \(x\in X\) and \(0\le t\le a\). We define \(g_1\in C_{T,h}\left( X, L(\mathbb {R})\right) \) as before. Namely, for \(x\in X\),

  • For each \(s\in H(x)\), we set

    $$\begin{aligned} g_1(x)(t) = f_1(x)(t) + \tilde{q}(T_s x)\cdot u(T_s x)(t-s) \quad \text {for}~t\in [s,s+a]. \end{aligned}$$
  • For \(t\in \mathbb {R}{\setminus } \bigcup _{s\in H(x)} [s,s+a]\), we set \(g_1(x)(t) = f_1(x)(t)\).

This satisfies \(|g_1(x)(t)-f(x)(t)| \le |g_1(x)(t)-f_1(x)(t)| + |f_1(x)(t)-f(x)(t)| \le 3\delta \).

We would like to show \(g_1(B)\cap g_1(C) = \emptyset \). Suppose \(x\in B\) and \(y\in C\) satisfy \(g_1(x)=g_1(y)\). There exist \(|s_1|\le a/4\) and \(|s_2|\le a/4\) satisfying \(T_{s_1} x\in B_0\) and \(T_{s_2} y\in C_0\). We can assume \(s_1\le s_2\) without loss of generality. Set \(\varepsilon =s_2-s_1 \in [0,a/2]\). We have

$$\begin{aligned} g_1(x)(s_1+t) = g(T_{s_1} x)(t) \text { and } g_1(y)(s_2+t) = g(T_{s_2} y)(t) \quad \text {for}~t\in [0,a]. \end{aligned}$$

\(g_1(x)=g_1(y)\) implies that

$$\begin{aligned} g(T_{s_1} x) (t+\varepsilon ) = g(T_{s_2} y) (t) \quad \text {for}~t\in [0, a-\varepsilon ]. \end{aligned}$$

It follows from the property (3) of g that \(d(T_{s_1}x, T_{s_2}y) < \delta \). Since \(\delta < d(B_0, C_0) \le d(T_{s_1}x, T_{s_2}y)\), this is a contradiction. Therefore \(g_1(B)\cap g_1(C) = \emptyset \). This proves the lemma. \(\square \)

Now we can prove Theorem 1.3. Note that X and \(X\times X\) are hereditarily Lindelöf (that means that every open cover of a subspace has a countable subcover). Using these facts and applying Lemma 3.3 to each point in \(X{\setminus } F\) and Lemma 3.4 to every pair of distinct points in \(X{\setminus } F\), there exist countable families of closed sets \(\{A_n\}_{n=1}^\infty \), \(\{B_n\}_{n=1}^\infty \) and \(\{C_n\}_{n=1}^\infty \) of \(X{\setminus } F\) such that

  • \(X{\setminus } F = \bigcup _{n=1}^\infty A_n\) and \((X{\setminus } F)\times (X{\setminus } F) {\setminus }\{(x,x):x\in X\} = \bigcup _{n=1}^\infty B_n\times C_n\).

  • \(G(A_n)\) are open and dense in the space \(C_{T,h}\left( X, L(\mathbb {R})\right) \) for all \(n\ge 1\).

  • \(G(B_n,C_n)\) are open and dense in the space \(C_{T,h}\left( X, L(\mathbb {R})\right) \) for all \(n\ge 1\).

By the Baire category theorem, the set

$$\begin{aligned} \bigcap _{n=1}^\infty G(A_n) \cap \bigcap _{n=1}^\infty G(B_n,C_n) \end{aligned}$$

is dense and \(G_\delta \) in \(C_{T,h}\left( X, L(\mathbb {R})\right) \). In particular it is not empty. Any element f in this set gives an embedding of the flow (XT) in \(L(\mathbb {R})\).

Remark 3.5

The proof of the Bebutov–Kakutani theorem in [1, 4] used the idea of “constructing large derivative”. It is possible to prove Theorem 1.3 by adapting this idea to the setting of one-Lipschitz functions. But this approach seems a bit tricky and less flexible than the proof given above. The above proof possibly has a wider applicability to different situations (e.g. other function spaces).